Anda di halaman 1dari 26

1597

Multi-zone modelling of partially premixed


low-temperature combustion in pilot-ignited
natural-gas engines
S R Krishnan* and K K Srinivasan
Department of Mechanical Engineering, Mississippi State University, Starkville, Mississippi, USA

The manuscript was received on 25 December 2009 and was accepted after revision for publication on 29 June 2010.

DOI: 10.1243/09544070JAUTO1472

Abstract: Detailed results from a multi-zone phenomenological simulation of partially pre-


mixed advanced-injection low-pilot-ignited natural-gas low-temperature combustion are
presented with a focus on early injection timings (the beginning of (pilot) injection (BOI))
and very small diesel quantities (2–3 per cent of total fuel energy). Combining several aspects of
diesel and spark ignition engine combustion models, the closed-cycle simulation accounted
for diesel autoignition, diesel spray combustion, and natural-gas combustion by premixed
turbulent flame propagation. The cylinder contents were divided into an unburned zone,
several pilot fuel zones (or ‘packets’) that modelled diesel evaporation and ignition, a flame
zone for natural-gas combustion, and a burned zone. The simulation predicted the onset of
ignition, cylinder pressures, and heat release rate profiles satisfactorily over a wide range
of BOIs (20–60u before top dead centre (before TDC)) but especially well at early BOIs. Strong
coupling was observed between pilot spray combustion in the packets and premixed turbulent
combustion in the flame zone and, therefore, the number of ignition centres (packets)
profoundly affected flame combustion. The highest local peak temperatures (greater than
2000 K) were observed in the packets, while the flame zone was much cooler (about 1650 K),
indicating that pilot diesel spray combustion is probably the dominant source of engine-out
emissions of nitrogen oxide (NOx). Further, the 60u before TDC BOI yielded the lowest average
peak packet temperatures (about 1720 K) compared with the 20u before TDC BOI (about 2480 K)
and 40u before TDC BOI (about 2700 K). These trends support experimental NOx trends, which
showed the lowest NOx emissions for the 60u, 20u, and 40u before TDC BOIs in that order.
Parametric studies showed that increasing the intake charge temperature, pilot quantity, and
natural-gas equivalence ratio all led to higher peak heat release rates and hotter packets but the
pilot quantity and intake temperature affected the potential for NOx formation to a greater extent.

Keywords: phenomenological model, low-temperature combustion, dual-fuel engines, pilot


ignition, micropilot, natural gas, Shell ignition model

1 INTRODUCTION natural-gas prices, natural-gas-fuelled engines re-


main commercially viable for several applications:
The continuing drive to adopt clean alternative fuels transportation, distributed electric power genera-
and to reduce dependence on conventional petro- tion, gas production, gas pipeline compression, and
leum fuels has stimulated the substitution of natu- petroleum-refining operations. However, the chal-
ral gas (among other fuels) for gasoline and diesel lenge with employing natural gas as an engine fuel
in internal combustion engines. Despite fluctuating stems from the fact that it is primarily composed of
methane, which is very difficult to ignite. Several
*Corresponding author: Department of Mechanical Engineering, ignition strategies including lean-burn spark ignition
Mississippi State University, PO Box 9552, Mississippi State, MS [1], laser ignition [2, 3], and pilot ignition [4–8] have
39762, USA. been utilized to increase efficiencies and to decrease
email: krishnan@me.msstate.edu emissions from natural-gas engines. Of these, pilot

JAUTO1472 Proc. IMechE Vol. 224 Part D: J. Automobile Engineering


1598 S R Krishnan and K K Srinivasan

ignition, which uses small diesel sprays as the ignition control of the combustion process, other LTC con-
sources, offers the advantages of multiple ignition cepts utilize early or late direct fuel injection.
centres and higher energy densities compared with Although early- and late-injection LTC concepts
conventional spark ignition. Gebert et al. [9] showed have their own advantages and drawbacks, the key
that the energy associated with a pilot diesel spray idea in both approaches is to ensure that the end of
volume of 1 mm3 is approximately two orders of fuel injection (EOI) is separated from the start of
magnitude greater than the energy provided by a spark. combustion (SOC) by ‘sufficiently long’ ignition de-
Therefore, the quality of ignition is better with diesel lay periods [17]. By comparison, both fuel injection
pilot ignition, and spatially dispersed, multiple-ignition and combustion events occur simultaneously in
centres lead to faster burn rates compared to single- conventional diesel engines, resulting in high NOx
point ignition sources. and PM emissions. Consequently, this separation
In pilot-ignited natural-gas engines, the natural- between the EOI and the SOC is necessary to reap
gas–air mixture is inducted during the intake stroke the NOx and PM benefits of LTC.
and ignited towards the end of compression by small Kalghatgi and co-workers [14–16] achieved con-
diesel sprays. These dual-fuel engines yield low trolled LTC at relatively high loads (indicated mean
oxides of nitrogen (NOx) and particulate matter effective pressure (IMEP), approximately 15 bar)
(PM) emissions and fuel conversion efficiencies by resorting to partially premixed combustion with
comparable with conventional diesel engines. How- gasoline. They demonstrated that late injection of
ever, high unburned hydrocarbon (HC) emissions, gasoline yielded low NOx and PM emissions by in-
cycle-to-cycle variations at low loads, and tendency creasing ignition delays substantially compared with
to knock at high loads are known problems in dual- diesel fuels. Since diesel is easier to autoignite than
fuel engines [6]. One approach for retaining the NOx gasoline, the maximum allowable engine loads in
benefits of dual-fuel engines is maximization of diesel-fuelled LTC HCCI are lower than for gasoline.
natural-gas substitution or minimization of diesel These results indicate that low-cetane fuels that are
usage [8]. To reduce NOx emissions further without highly resistant to autoignition (e.g. gasoline and
natural gas) can yield controlled LTC over a wide
compromising HC emissions or fuel conversion
range of engine operating conditions.
efficiency, more sophisticated diesel injection sys-
tems (than conventional mechanical injection sys-
tems) that can consistently and reliably meter very
2 ADVANCED-INJECTION LOW-PILOT-IGNITED
small diesel quantities are required [7]. Better pilot NATURAL-GAS COMBUSTION
injection systems will also enable dual-fuel low-
temperature combustion (LTC) strategies to be Dual–fuel experiments [18–20] performed with a
adopted. common-rail low-pilot-injection system established
For conventional spark ignition (SI) and compres- that engine-out NOx emissions can be reduced to
sion ignition engine combustion, Flynn et al. [10] about 0.2 g/kW h over a wide range of engine loads
concluded that engine-out NOx emissions are in- when very small diesel sprays (providing 2–3 per
evitable and may not be decreased below a limit cent of total fuel energy) are injected early in
because the local in-cylinder temperatures are usu- the compression stroke (about 60u before top dead
ally higher than 2000 K. Therefore, to meet future centre (before TDC)) to ignite premixed natural-gas–
emissions standards and to increase fuel conversion air mixtures. The penalty on fuel conversion effi-
efficiencies simultaneously, LTC concepts that can ciencies was minimal despite the remarkable reduc-
provide controlled engine operation at high engine tion in NOx emissions; however, HC and carbon
loads (brake mean effective pressure (BMEP), greater monoxide (CO) emissions increased substantially at
than 10 bar) are desirable. Different approaches have advanced injection timings. This partially premixed
been investigated to realize LTC operation with LTC concept has been termed advanced-injection
gasoline and diesel fuels. These include, among low-pilot-ignited natural-gas (ALPING) combustion
others, homogeneous charge compression ignition [20]. By combining a low-cetane main fuel (natural
(HCCI) [11], early-injection LTC (e.g. the UNIBUS gas) and early injection of a high-cetane pilot fuel
concept [12]), late-injection LTC with aggressive (diesel), the EOI and the SOC were separated to
exhaust gas recirculation (EGR) and swirl ratios [13], achieve controlled LTC for a BMEP of up to 12 bar.
and partially premixed LTC [14–16] with low-cetane It is hypothesized that the ALPING combustion
fuels. Combustion control issues with HCCI limit process combines some aspects of both conven-
the maximum attainable BMEPs. To provide better tional diesel and SI engine combustion. Early in-

Proc. IMechE Vol. 224 Part D: J. Automobile Engineering JAUTO1472


Multi-zone modelling of partially premixed low-temperature combustion 1599

jection (the beginning of (pilot) injection (BOI)) of 3 bar. Also, the earliest possible BOI was extended
pilot diesel leads to a very long ignition delay period from 60u before TDC without EGR to 70u before TDC
because the local in-cylinder temperatures at BOI with hot EGR.
are not conducive to ignition. Over the long ignition
delay period, the pilot sprays disperse and mix with
the surrounding natural-gas–air mixture. When the
3 OBJECTIVES
piston approaches top dead centre (TDC) and the
local temperatures become high enough to initiate
A phenomenological simulation that provides a
and sustain rapid pre-ignition reactions, the diesel
framework for predicting dual-fuel, partially pre-
autoignites (simultaneously or in rapid succession)
mixed LTC can be employed to gain deeper insights
at different locations inside the cylinder. These
into ALPING combustion, especially at early BOIs,
spatially distributed ignition centres foster the
which yielded very low NOx emissions. Phenomen-
creation of several ‘flamelets’ that result in localized
ological combustion models are routinely used to
flame propagation through the lean mixture of
simulate performance and emissions in diesel
natural gas (mostly methane) and air. These dis-
engines [24–26]. For simulating dual-fuel engine
tributed flamelets ensure large enflamed areas, thus
combustion, several models have been developed.
resulting in faster energy release rates and good fuel
They range from relatively simple two-zone models
conversion efficiencies.
[27] to increasingly complex multi-zone models [28]
The range of equivalence ratio (w) values (based on
(some of which include detailed chemical kinetics
both diesel and natural gas) encountered in ALPING
[29–31]), and multi-dimensional models [32]. These
combustion lies between 0.2 (for low loads) and 0.6
models predict conventional dual-fuel engine per-
(for high loads). Whereas it was previously thought
formance and emissions satisfactorily. However, as
that flame propagation in methane–air mixtures was
discussed above, ALPING combustion differs sig-
limited to a lean flammability limit of w < 0.5 for
nificantly from conventional dual-fuel engine com-
atmospheric flames, recent numerical simulations
bustion. A review of the literature revealed that no
[21] suggested that, once ignition is achieved, flame
existing models have been developed to simulate
propagation may occur even at much leaner equiva-
dual-fuel, partially premixed LTC explicitly at very
lence ratios (w < 0.2) under engine-like conditions of
early BOIs. Consequently, the specific objectives of
temperature and pressure. This evidence supports the
the present work are as follows:
hypothesis that flame propagation may occur from
multiple ignition centres in ALPING combustion. (a) to develop a multi-zone phenomenological
The research of Kalghatgi and co-workers [14–16] simulation of ALPING combustion and to
emphasizes the performance and emissions benefits predict the cylinder pressure and heat release
of partially premixed LTC strategies. The ALPING histories over a range of BOIs, with specific
combustion process is a partially premixed LTC con- emphasis on understanding combustion at
cept realized with a specific combination of two early BOIs (50–60u before TDC);
fuels: an easily ignitable fuel (e.g. diesel) is used to (b) to use the ALPING combustion simulation to
initiate lean premixed combustion of a relatively investigate the effects of the intake manifold
difficult-to-ignite fuel (e.g. natural gas). Since diesel temperature, pilot fuel quantity, and natural-gas
is just sufficiently (but not completely) premixed equivalence ratios for the early BOI of 60u before
with the natural-gas–air mixture, ALPING combus- TDC.
tion occurs at locally lean equivalence ratios for early
BOIs, thereby leading to lower local temperatures 4 MODEL DEVELOPMENT
and very low NOx emissions. Further, with lean
premixed ALPING combustion, PM emissions are A multi-zone phenomenological combustion model
also expected to be insignificant compared with that simulates closed-cycle engine processes has
conventional diesel combustion. High HC and CO been developed. The primary objective of the model
emissions and poor low-load engine stability with is to simulate early-BOI ALPING combustion, which
ALPING combustion were addressed in hot EGR is significantly more difficult to predict than re-
experiments [22, 23], which yielded 70 per cent HC tarded-BOI combustion. In ALPING combustion, the
emissions reduction, efficiency improvement by 5 initial pilot combustion phase is somewhat similar
percentage points, and more stable engine operation to diesel spray combustion while the subsequent
with virtually no NOx penalty at BMEPs of 6 bar and natural-gas combustion phase involves premixed

JAUTO1472 Proc. IMechE Vol. 224 Part D: J. Automobile Engineering


1600 S R Krishnan and K K Srinivasan

turbulent combustion by flame propagation (simi- and energy added owing to combustion in the
lar to an SI engine). Therefore, the present model packets or the flame zone. The second and third
incorporates phenomenological aspects of both terms represent the boundary work output and
diesel spray combustion and premixed turbulent the net enthalpy input due to any mass transfer
flame propagation. respectively. Of these energy interactions, heat trans-
The simulation starts from intake valve closure fer and boundary work are experienced by all zones.
(IVC) and proceeds until exhaust valve opening. Specific details of energy and mass interactions,
Depending on the stage of the simulation, the together with the submodels used in different zones,
cylinder contents are divided into one or more of are discussed later.
the following zones: an unburned zone, pilot fuel
zones (packets), a flame zone, and a burned zone, as
shown schematically in Fig. 1. While the different 4.1 Model assumptions
zones are illustrated as continuous and contiguous
The important assumptions of the phenomenologi-
for ease of visualization, it should not be construed
cal ALPING combustion simulation are listed below.
that they are necessarily so; in fact, at advanced
BOIs, packets can be dispersed throughout the 1. Each zone is treated as an open thermodynamic
cylinder, forming distributed ignition centres. Since system consisting of a mixture of ideal gases.
the simulation is only quasi-dimensional and is 2. Heat transfer occurs only between each zone and
not capable of providing spatial resolution of the the cylinder walls. Although mass transfer is
cylinder contents, it does not require any of these allowed between zones, interzonal heat transfer
zones to be continuous or contiguous. is neglected.
The generic form of the energy equation for a 3. Since the natural gas used in ALPING combustion
given zone i (an open thermodynamic system) is experiments was largely composed of methane
(98 vol %), methane is assumed to represent
dðmi cvi Ti Þ X dQi P dVi X dmi
~ { z hi ð1Þ natural gas in the simulation and n-dodecane is
dh net in
dh dh net in
dh assumed to represent diesel fuel.
4. When diesel and natural gas burn in any zone,
where mi is the mass, cvi is the specific heat at they are converted into products of complete
constant volume, Ti is the temperature, and all combustion (carbon dioxide (CO2) and water
derivatives are with respect to the crank angle (CA) h. (H2O)).
The first term on the right-hand side of equation (1) 5. The end of combustion (EOC) is attained when
is the net heat input into the zone, which includes the unburned zone mass becomes very small (less
heat transfer from and to the cylinder walls and, if than 0.001 per cent of its initial value) or more
applicable, energy lost owing to diesel evaporation than 99.9 per cent of the total entrained mass in
the flame zone has been burned.

4.2 Unburned zone


At IVC, the premixed natural-gas-air mixture trapped
in the cylinder constitutes the unburned zone. Mass
is transferred from the unburned zone to packets
after the BOI and to the flame zone after the SOC.
The state of the unburned zone is tracked as long as
there is some unburned mass in the cylinder.

4.3 Pilot fuel zones (packets)


At the BOI, the packets, which account for combus-
tion of diesel and some natural gas entrained in the
diesel spray, come into existence. The lifetime of a
Fig. 1 Conceptual schematic diagram of zone evolu- packet depends on the fulfilment of certain ‘dump-
tion in the multi-zone simulation ing criteria’ that are discussed later (see the descrip-

Proc. IMechE Vol. 224 Part D: J. Automobile Engineering JAUTO1472


Multi-zone modelling of partially premixed low-temperature combustion 1601

tion of the burned zone). A user-specified number of taken (in degrees CA) for the spray to penetrate a
packets is sequentially injected throughout the characteristic entrainment distance equal to the
injection duration and an axisymmetric diesel spray cylinder bore radius. From a physical perspective,
is assumed. As shown in Fig. 2, packets are classified hch can be interpreted as the time taken to entrain
on the basis of their time of entry into the cylinder (I) the surrounding fluid completely into the diesel
and the anticipated radial stratification in unburned spray or the time taken for the diesel spray to engulf
mixture entrainment across the spray (J). In other the entire combustion chamber volume. To calculate
words, J packets of equal diesel mass are injected hch, spray penetration is determined using the
every time step over the duration of injection, correlation recommended by Dent [33]. While other
leading to a constant rate of fuel injection. A higher more sophisticated spray penetration models (e.g.
value of I indicates that the packet is injected the Siebers [34] model) are certainly available, the
relatively later in the injection process, while a Dent model has been adopted in the present work
higher value of J indicates that the packet’s location because it is more straightforward to implement in a
is closer to the spray axis and less capable of phenomenological context. In any case, the spray
entraining unburned mixture. Although spatial pac- penetration model used in the current entrainment
ket locations are not determined in this model, this model only affects overall entrainment indirectly (via
method of packet identification allows appropriate hch) compared with the entrainment constant K,
stratification of mixture entrainment across the spray, which is an important model parameter, as dis-
and, in this sense, the entrainment model is quasi- cussed below.
dimensional. In the following paragraphs, various From the aforementioned scaling arguments, the
submodels relevant to the packets including mixture total rate dmtot-ent/dh of unburned mass entrained
entrainment, diesel evaporation, ignition, and diesel into the spray must be directly proportional to the
and natural-gas combustion are discussed. unburned-zone mass and inversely proportional to
the characteristic entrainment time and can be
written as
4.3.1 Mixture entrainment model
dmtot-ent mu
The phenomenological basis of the entrainment ! ð2Þ
dh hch
model is the recognition that the mixture entrain-
ment into the spray is simultaneously governed by Since the total unburned entrained mass is distrib-
several factors. Since the spray can entrain unburned uted among a total number of Itot 6 Jtot packets,
mixture only if it is available, the entrainment rate dmtot-ent/dh is divided by Itot 6 Jtot. Also, proper
is proportional to the unburned-zone mass mu. stratification in mixture entrainment must be en-
Further, all the available unburned mixture can be sured in the direction normal to the spray axis and
entrained into the spray in a characteristic entrain- for packets injected later in the injection process.
ment time hch, which is assumed to be the time Finally, the decay of spray entrainment as the spray

Fig. 2 Schematic diagram of packets classification showing spray entrainment and radial
stratification

JAUTO1472 Proc. IMechE Vol. 224 Part D: J. Automobile Engineering


1602 S R Krishnan and K K Srinivasan

disintegrates and loses its initial momentum must be BOIs in ALPING combustion [37, 38]. Therefore, the
captured. Based on these considerations and follow- Shell autoignition model [39] was used to model
ing Bell [35], the mass ment of mixture entrained into ignition in each packet. The Shell model employs
any packet (I, J) in one calculation step Dh is generic species in a global eight-step chain-branch-
expressed as ing reaction mechanism of hydrocarbon ignition and
  includes the reactions for chain initiation, propaga-
mu Dh I J tion, branching, and termination. The original model
ment ðI, J Þ~K exp { ð3Þ
Itot Jtot ð1zY Þ hch Itot Jtot coefficients for fuel with a research octane number
(RON) of 90 (assuming n-dodecane stoichiometry)
where K is the entrainment constant, which is an were used with the exception of the pre-exponential
important model parameter that influences the net factor Ap3 controlling fuel consumption in the
spray momentum, Y is the ratio of time elapsed since propagation cycle, which was modified from the
the BOI to the total injection duration, and Itot and original value of 161013 to 861011 to provide better
Jtot are the maximum values of I and J respectively. ignition delay predictions [37].
The exponential term in equation (3) accounts for Pre-ignition energy release, which is predicted in
stratification of entrainment across the spray and the ignition model, raises the packet temperatures,
progressively lower entrainment in packets that are and ignition of a packet is assumed to occur when
injected later in the injection process (higher I either the packet temperature exceeds 1100 K or the
values). The term 1 + Y ensures that the overall rate rate of increase in the packet temperature exceeds
of mixture entrainment in the spray decreases with 107 K/s. These two conditions have been used by
increasing elapsed time since the BOI. This mimics several researchers, including Halstead et al. [39]. In
reality because, as the spray penetrates into the addition to these ignition criteria, it is reasonable to
surroundings, its velocity decreases and eventually expect ignition to occur only if the equivalence ratio
mixture entrainment into the spray decreases. in each packet is within ignition limits. As observed by
Sazhina et al. [40], artificial enforcement of ignition
limits is necessary since the Shell model does not
4.3.2 Diesel evaporation model explicitly account for such limits. Therefore, in the
The total amount of pilot diesel fuel is initially present work, a packet is assumed to ignite if it
divided equally between all the packets, which are satisfies either of the aforementioned temperature
tracked separately thereafter. However, only evapo- criteria and its equivalence ratio is within the
rated diesel is allowed to enter the packets. All fuel assumed rich ignition and lean ignition limits of
droplets are initialized with the same initial user- w 5 3 and w 5 0.2 respectively. Upon ignition, the Shell
specified Sauter mean diameter (SMD). Following ignition model is deactivated for the packet and the
Kanury [36], the droplet evaporation rates are cal- packet combustion model is triggered.
culated using a quasi-steady isolated droplet eva-
poration model in which the evaporation rates
adjust instantly to the temperature and species con- 4.3.4 Packet combustion model
centration changes. At the end of each time step, the After a packet ignites, diesel is allowed to burn ac-
droplet diameters in each packet are updated on the cording to the single-step global reaction mechan-
basis of the remaining liquid volumes. ism adapted from Westbrook and Dryer [41], which
is given by

4.3.3 Ignition model  


{EA 1:5 Vij MW d
In pilot-ignited natural-gas engines, diesel fuel ig-
_ d-ij ~Apd exp
m ½C12 H26 0:25
ij ½O2 ij ð4Þ
RTij v
nites first and then initiates natural-gas combustion.
In this model, diesel ignition is modelled in the
packets. Although natural gas may take part in the where m _ d-ij is the diesel reaction rate (g/rad) in
diesel ignition process [6], its specific effects are not packet (I, J), the pre-exponential factor Apd is as-
very well known and, hence, are not considered here. sumed to be 2.561010 (mol/cm3)20.75/s (following
Empirical Arrhenius-type ignition delay correlations, trends in reference [41], this is lower than the value
which were originally developed for predicting the of 3.861010 recommended for decane), the activa-
ignition delays in conventional diesel combustion, tion energy EA is 30 kcal/mol, [C12H26]ij is the packet
greatly underpredicted the ignition delays for early diesel (n-dodecane) vapour concentration (mol/cm3),

Proc. IMechE Vol. 224 Part D: J. Automobile Engineering JAUTO1472


Multi-zone modelling of partially premixed low-temperature combustion 1603

[O2]ij is the packet oxygen concentration (mol/  


cm3), Tij is the packet temperature (K), Vij is {EA
_ ng-ij ~Apng exp
m ½CH4 {0:3
ij ½O2 1:3
ij
the instantaneous packet volume (cm3), R is the RTij
universal gas constant (kcal/mol K), MWd is the    Cw 
Vij MW ng  
molecular weight of n-dodecane, and v is the | 1{1{wij  ð7Þ
v
crankshaft angular velocity (rad/s). From equation
(4), it is obvious that, for a given packet temperature, where wij and Cw have the same values as in equation
the diesel reaction rate will increase when either the (5).
fuel or oxygen concentration is increased. As the BOI The actual diesel and natural-gas burn rates are
is advanced from 20u before TDC to 60u before TDC, taken to be the lower of two rates: the rates at which
progressively higher levels of oxygen entrainment in they become available (by evaporation or entrain-
packets are allowed because of longer ignition ment) in the packet or the burn rates predicted by
delays, thereby increasing the reaction rates and equations (4) to (7). The heat release rates in all
heat release rates without limits. In contrast with the ignited packets are determined as the product of the
behaviour of equation (4), the experimental peak actual fuel mass burning rates and their respective
heat release rates at advanced BOIs were found to be lower heating values.
lower than at intermediate BOIs (e.g. 35–40u before
TDC). This may be expected because the diesel pilot,
when allowed to mix with the lean natural-gas–air 4.4 Flame zone
mixture over a long ignition delay period, will
The flame zone, which models flame propagation in
probably be very lean in some spatial locations,
the natural-gas–air mixture, is formed when ignition
thereby decreasing the local reaction rates. To
first occurs in any packet and ceases to exist at the
capture these trends, equation (4) was modified
EOC. Conceptually, the flame zone refers to the
only for lean packet equivalence ratios as
combustion zones surrounding the ignited packets.
  Since spatially distributed combustion zones cannot
{EA be handled separately in a quasi-dimensional com-
_ d-ij ~Apd exp
m ½C12 H26 0:25 1:5
ij ½O2 ij
RTij bustion model, these zones are lumped into a single
  Cw  ‘flame zone’. Unlike traditional SI engine combus-
Vij MW d  
| 1{1{wij  ð5Þ tion, ALPING combustion does not originate from a
v
single location inside the cylinder. Consequently, the
computation of the flame area is challenging, and
where wij (( 1) is the total equivalence ratio of packet geometric modelling of the flame area as one
(I, J) considering both diesel and natural gas, and the spherical flame front (cf. reference [42]) with a
lean packet equivalence ratio exponent Cw is set single point of origin (e.g. a spark plug) is neither
realistic nor feasible. Therefore, the flame area is
equal to 0.2.
treated in a rather unique way (based on the packet
Similar to packet diesel combustion, natural gas
areas) in this paper, as explained later.
(methane) entrained into each packet together with
To model natural-gas combustion by premixed
air is assumed to burn according to the single-step
turbulent flame propagation, the entrainment and
global reaction mechanism [41]
burn-up model of Tabaczynski et al. [43] is adopted.
  In this model, the small-scale turbulence structure
{EA model proposed by Tennekes [44] is used. Small-
_ ng-ij ~Apng exp
m ½CH4 {0:3
ij ½O2 1:3
ij
RTij scale turbulence is described as ‘vortex tubes’ of
diameter g (the Kolmogorov microscale) that are
Vij MW ng
| ð6Þ stretched by eddies of size l (the Taylor scale). The
v
main combustion implications of the Tennekes
model are slow burning at the laminar burning
where m _ ng-ij is the natural gas reaction rate (g/rad) in velocity in the l-scale regions and fast burning in the
packet (I, J), Apng 5 8.36105 s21, and EA 5 30 kcal/ g-scale regions. The Tennekes model has previously
mol. Also, to be consistent with the diesel reaction been implemented and validated by Tabaczynski
rate expression given in equation (5), the natural-gas et al. [43] and further refined by Tabaczynski et al.
reaction rate was modified only for lean total packet [45] to simulate turbulent flame propagation in SI
equivalence ratios as engines. With this model, the mass entrainment rate

JAUTO1472 Proc. IMechE Vol. 224 Part D: J. Automobile Engineering


1604 S R Krishnan and K K Srinivasan

_ e in the flame zone can be expressed as


m In equation (10), the proportionality constant KAf
(set to 0.9) and the burned mass fraction exponent
_ e ~ðru rf Þ1=2 Af ue
m ð8Þ Cb (set to 0.4) are model parameters. The number of
ignited packets at any CA is npign, the packet burned
where ru is the unburned-zone density, rf is the mass fraction is ypb, and the packet volume is Vp.
flame-zone density, Af is the instantaneous mean Based on dimensional considerations, the packet
flame area, and ue is the turbulent entrainment area may be related to its volume raised to the two-
velocity. Since equation (8) does not have an explicit thirds power; this relationship between the area and
dependence of the entrainment rate on the available the volume has been used previously for heat
unburned mass, flame entrainment remains inordi- transfer area calculations by Bell [35] and Krishnan
nately large even near the EOC when the unburned- [46].
zone mass becomes very small. To overcome this The turbulent entrainment velocity ue, which can
issue, equation (8) was modified to account expli- be interpreted as the speed of propagation of
citly for the reduced availability of unburned mass ignition centres into the unburned gas, is deter-
towards the EOC, and the mass entrainment rate in mined from the equation
the flame was calculated as  1=2   
   ru {Rf
{mu ue ~SL zC1 u’ 1{exp ð11Þ
_ e ~ðru rf Þ1=2 Af ue 1{exp
m ð9Þ rf C2 L
cfdrop mtot
where SL is the laminar burning velocity evaluated at
where the exponential term ensures that the en- instantaneous in-cylinder conditions, u9 is the
trainment rate decreases exponentially as the un- turbulence intensity, Rf is the equivalent spherical
burned mass fraction mu/mtot in the cylinder radius of the flame calculated from Af, and L is the
decreases below a critical value that is dependent integral length scale. To find ue, SL is determined
on the model constant cfdrop. The value of cfdrop was using a typical correlation for methane–air mixtures
set to 0.0625 to ensure that the entrainment rate [47]. The values of C1 and C2 are 4 and 1 respectively.
tapers off smoothly when more than 90 per cent of The exponential term in equation (11), which has
the unburned mixture has been entrained. been used before in SI combustion modelling [48,
To evaluate the flame entrainment rate from 49], has been included to account explicitly for the
equation (9), both Af and ue need to be determined. increased importance of the turbulence intensity
In ALPING combustion, several non-stationary dis- (relative to the laminar burning velocity) with
tributed ignition centres are probable, resulting in increasing flame size. The initial value of L at the
many localized flamelets that may eventually inter- SOC is proportional to the instantaneous combus-
act with each other to form a single turbulent flame tion chamber height with the proportionality con-
or may retain their individuality throughout the stant C4 set to 0.2, following Hill [50]. The initial
combustion process. Therefore, Af should be calcu- value of u9 at the SOC is assumed to be proportional
lated while accounting for the possible presence of to the mean piston speed with the proportionality
multiple ignition centres. Also, increasing the num- constant Cu9 equal to 0.6. The values of L and u9 at
ber of ignition centres should lead to a larger any subsequent time step are determined by
enflamed area and faster combustion rates. Accord- assuming angular momentum conservation for in-
ingly, the instantaneous mean flame area is assumed dividual eddies during rapid compression or expan-
to be the envelope of the instantaneous enflamed sion (rapid distortion theory) following reference
area of ignited packets. For simplicity, each packet is [43], according to
presumed to retain its individuality throughout the  1=3
combustion process, i.e. no interaction, which may r
L~Li ui ð12Þ
lead to overlapping or combining of enflamed areas ru
is allowed between the packets. The instantaneous
flame area is then determined from the enflamed
 1=3
area of the ignited packets as r
u’~u’i u ð13Þ
rui
np ign
X Cb 2=3
Af ~KAf ypb Vp ð10Þ
where the subscript i refers to the initial values of the
n~1
respective variables.

Proc. IMechE Vol. 224 Part D: J. Automobile Engineering JAUTO1472


Multi-zone modelling of partially premixed low-temperature combustion 1605

The turbulent mass burn rate m_ b is proportional to ‘dumping’) is expected to capture hot regions of
the total unburned mass (total entrained mass – total burned products in packets, which lead to most of
burned mass) in the flame zone and is determined the NOx formed during combustion. When all
from the equation packets are dumped, the flame area becomes zero,
the flame zone ceases to exist, and combustion is
metot {mbtot terminated.
_ b~
m zru Af SL ð14Þ
tb

where metot is the cumulative mass entrained into 4.6 Heat transfer model
the flame zone, mbtot is the cumulative mass burned
The overall heat transfer rate dQtot/dh from cylinder
in the flame zone, and tb is the characteristic burn
gases to cylinder walls is determined using the
time, which is determined as
Woschni [53] heat transfer correlation and the
C3 l average cylinder temperature Tavg. Following refer-
tb ~ ð15Þ ence [35], the heat transfer rate dQi/dh from each
SL
zone i is obtained by weighting dQtot/dh with the
The burn time represents the characteristic time area (computed as volume raised to the two-thirds
taken for laminar burning to occur in a length scale power) and temperature Ti of that zone according to
corresponding to the Taylor scale l. The constant C3 2=3
was assigned a value of 2.5 following reference [48]. dQi dQtot Vi ðTi {Twall Þ
~ P 2=3 ð17Þ
Assuming isotropic turbulence, l is calculated from dh dh V ðTi {Twall Þ
i i
the expression [51]
 1=2  {1=2 where Vi is the zone volume and Twall is the cylinder
l 15 u’L wall temperature (set equal to 480 K).
~ ð16Þ
L A n

where the constant A is set equal to 1, n (5 m/rf) is the 5 RESULTS AND DISCUSSION
kinematic viscosity, and the dynamic viscosity m is
determined from the expression m 5 3.361027 Tf0:7 In this section, predictions from the phenomenolo-
given in reference [52]. gical simulation are compared with experimental
results at different BOIs for ALPING combustion.
The flame zone heat release rate is evaluated as
The engine details and operating conditions (used as
the product of the mass burn rate and the lower
model inputs) are given in Table 1. A 2.4 l single-
heating value of methane. The mass burned in the
cylinder research engine equipped with a dedicated
flame zone in one time step is transferred to the
pilot injection system was used for the ALPING
burned zone in the subsequent time step.
experiments. The diesel injected quantity was fixed
at 3.3 g/min (the minimum value possible with the
injection system) for all BOIs to ensure minimal NOx
4.5 Burned zone
The burned zone is formed one time step after the Table 1 Engine details and operating conditions
SOC and contains burned products transferred every Parameter Specification
time step from the flame zone. The burned products Engine type Single cylinder, four stroke
in a packet are not transferred to the burned zone Bore (mm) 137
until the following conditions are satisfied: first, the Stroke (mm) 165
Compression ratio 14.5
packet temperature is lower than a critical tempera- Combustion system Direct injection, Mexican hat
ture (1900 K), below which NOx formation rates will Diesel injection system Electronic, common rail
Diesel injector Pencil type, four holes
become negligible, and, second, either more than 90 BOI Variable (20–60u before TDC)
per cent of diesel and natural gas in a packet have Diesel injection pressure (bar) 359.5
burned or their reaction rates have become less than Diesel injected quantity (g/min) 3.3
Diesel injection duration (deg CA) 5
1 per cent of their maximum values. Natural-gas fuelling rate (g/min) 77.3–94.0 (variable with BOI)
When both the first condition and the second Engine speed (r/min) 1700
Experimental BMEP associated 6
condition are satisfied, all packet contents are with fuelling rates (bar)
allowed to mix instantaneously and adiabatically Intake manifold pressure (kPa) 181
Intake manifold temperature (K) 348
with the burned zone. This mixing process (termed

JAUTO1472 Proc. IMechE Vol. 224 Part D: J. Automobile Engineering


1606 S R Krishnan and K K Srinivasan

emissions. However, natural-gas fuelling was varied mentally derived heat release rates are ‘differen-
to maintain a BMEP of 6 bar for all BOIs at a constant tiated’ quantities that are dependent either on local
speed of 1700 r/min. The intake manifold pressure packet conditions or on the pressure derivative,
and the temperature were fixed at 181 kPa and 348 K while cylinder pressures are ‘integrated global’
respectively. All the model parameters were cali- quantities. Therefore, while comparing the experi-
brated (see reference [38] for details) for a baseline mental and predicted heat release rates throughout
test case (the 40u before TDC BOI) and systematic this paper, it is important to remember that,
model sensitivity studies were performed to obtain although the cylinder pressures may be predicted
optimal values (Table 2). Subsequently, all model satisfactorily, a perfect match between the heat
parameters (except Jtot as discussed below) were held release rates may not exist. In fact, some differences
constant for all other experimental conditions (dif- may be expected in the heat release histories
ferent BOIs) discussed in this paper. The simulation because of the fundamental differences between
was first validated with experimental pressure and the combustion models used in the simulation and
heat release histories over the entire range of BOIs the experimental heat release analysis and their
from 20u before TDC to 60u before TDC. Then, model associated assumptions and limitations. The diffi-
predictions at the early BOI of 60u before TDC were culties in matching the experimental and simulated
examined closely. Finally, parametric studies were heat release and the accuracy to which the heat
performed at the early BOI of 60u before TDC with release histories can be predicted were also dis-
different intake charge temperatures, pilot fuel cussed by Hountalas et al. [54] for diesel engines.
quantities, and natural-gas equivalence ratios.

5.1.1 Ignition delay behaviour


5.1 Model validation at various BOIs In the early stages of model development, it was
Before discussing the model validation results, it is determined that predicting the onset of ignition over
important to clarify the extent to which model a wide range of BOIs is critically important to predict
predictions can be expected to match experimental ALPING combustion accurately. Two different types
results. While it is relatively straightforward to of ignition model were originally considered [38]:
predict the cylinder pressures (since the cylinder (a) Arrhenius-type ignition delay correlations (typi-
pressure is an average global quantity), the predicted cally used in conventional dual-fuel combustion
heat release rates are profoundly influenced by local models);
(packet) reaction rates. Since the experimental heat (b) the Shell autoignition model.
release rates are determined using a two-zone
combustion model [46], they are dependent on the The Arrhenius-type ignition delay correlations
first derivative of the measured cylinder pressure. were unable to predict the experimental ignition
Therefore, in this sense, both predicted and experi- delay trends satisfactorily and, consequently, the
Shell model was adopted. The reader is referred to
Table 2 Model parameter values reference [37] for a detailed discussion of the
Parameter Value
shortcomings of Arrhenius-type ignition delay cor-
relations and the need for the Shell model to predict
Computational time step Dh (deg CA) 1/32
Entrainment constant K 0.05
pilot ignition in ALPING combustion.
Pre-exponential factor Apd in diesel reaction rate 2.561010 Figure 3 shows the experimental and predicted
((mol/cm3)20.75/s) ignition delays (with the Shell model) for different
Lean packet equivalence ratio exponent Cw 0.2
Flame area proportionality constant KAf 0.9 BOIs. To determine the experimental ignition delays,
Turbulent entrainment velocity constant C1 4.0 the CA-resolved cylinder pressure histories were
Turbulent entrainment velocity constant C2 1.0
Characteristic burn time proportionality constant C3 2.5
measured using a Kistler model 7061B pressure
Integral length scale proportionality constant C4 0.2 transducer, and the experimental heat release rates
Flame zone entrainment drop constant Cfdrop 0.0625 were determined using the two-zone combustion
Turbulence intensity proportionality constant Cu9 0.6
Packet burned mass fraction exponent for flame 0.4 model [46]. The pilot injection timings (BOIs) were
area calculations Cb determined using CA-resolved measurements from
Total number Itot of I packets 41
Total number Jtot of J packets 3–12
an injection pressure sensor mounted on the fuel
Initial diesel droplet diameter (SMD) (mm) 30 injector body. The onset of ignition (SOC) was
Shell model constant Ap3 for a primary reference 861011 determined as the start of heat release (the CA when
fuel of RON 90
positive heat release rates were first observed), and

Proc. IMechE Vol. 224 Part D: J. Automobile Engineering JAUTO1472


Multi-zone modelling of partially premixed low-temperature combustion 1607

possible at early BOIs, Jtot is increased as the BOI is


advanced to account for radial stratification of
mixture entrainment.
Figure 4 shows the predicted (subscript sim) and
experimental (subscript expt) pressures and heat
release rates for different BOIs. For these results, all
model parameters were fixed at the values specified
in Table 2. The only parameter that was changed
with the BOI was Jtot, whose specific value at each
BOI is shown in the corresponding figure. In Fig. 4,
slight discrepancies may be noted between some of
the experimental and simulated pressure curves even
before ignition. The primary experimental causes
for these discrepancies were as follows:

(a) the long valve overlap period (about 45u CA) in


the single-cylinder engine and the relatively late
IVC (37u after bottom dead centre) that led to
Fig. 3 Experimental and Shell-model-predicted igni- inaccuracies in the calculation of the total mass
tion delays at different BOIs trapped inside the cylinder at IVC from the
measured intake manifold conditions or flow-
the experimental ignition delay was defined as the rates;
difference between the BOI and the SOC. Figure 3 (b) the location of the cylinder pressure transducer,
shows that the experimentally measured ignition which was mounted via connecting passages to
delay increases with progressively earlier pilot injec- the combustion chamber (not flush mounted),
tion. As the BOI is advanced, the in-cylinder owing to the lack of direct access ports in the
temperatures at the time of injection are substan- cylinder head of the single-cylinder engine.
tially lower, resulting in slower preignition reactions,
Although the model was calibrated [38] with mo-
longer ignition delay periods, and thus lower NOx
toring pressure traces, the best match still showed
emissions. As is evident from Fig. 3, the predicted
some discrepancies between measurements and
ignition delays match quite well the experimental
predictions. These same discrepancies were also ob-
values over the entire BOI range. It must be noted
served when the measured and predicted combus-
that these predictions were obtained with all original
tion pressure traces were compared.
Shell model constants for a primary reference fuel of
For the late BOI of 20u before TDC, Fig. 4(a) shows
RON 90 except Ap3, which was modified to 861011
that the predicted pressures match the measured
[37].
pressures quite well but the peak heat release rate is
underpredicted. While the SOC is predicted well, the
5.1.2 Heat release and combustion pressure histories peak heat release rates are underpredicted and the
combustion rates are slower. The reason for slower
To achieve good combustion pressure and heat combustion at the 20u before TDC BOI is evident
release predictions at all BOIs, many model para- when the packet equivalence ratios are examined
meters were simultaneously optimized during the closely. Figure 5 shows the packet equivalence ratio
calibration phase [38] to obtain the values given in distributions at the SOC in terms of the percentage
Table 2. Sensitivity studies showed a strong coupling of packets that fall within a given range of equiva-
between combustion in packets and flame combus- lence ratios for five different BOIs. This representa-
tion. As the BOI is advanced, the injected diesel fuel tion (in terms of percentages) is necessary to compare
will undergo better mixing over longer ignition delay the packet equivalence ratio distributions over the
periods and a greater number of ignition centres will range of BOIs that use different total numbers of
be available throughout the combustion process. packets. In Fig. 5, a given packet equivalence ratio
Based on this premise, it was decided that the total range value indicates the range between that value
number of ignition centres (packets) should be and the immediately higher equivalence ratio value;
increased with BOI advancement. Since greater for example, an equivalence ratio range value of 0.6
diesel spray penetration and radial stratification are signifies the equivalence ratio range 0.6–0.7.

JAUTO1472 Proc. IMechE Vol. 224 Part D: J. Automobile Engineering


1608 S R Krishnan and K K Srinivasan

Fig. 4 Comparison of the experimental and predicted heat release rates (HRRs) and cylinder
pressures P for different BOIs

Proc. IMechE Vol. 224 Part D: J. Automobile Engineering JAUTO1472


Multi-zone modelling of partially premixed low-temperature combustion 1609

overpredicted. From Fig. 5, it may be observed that


almost 70 per cent of the packets lie within the
equivalence ratio range 0.8–1.1 for 40u before TDC.
Therefore, the combustion rates are faster and the
packet temperatures are also the highest for 40u
before TDC. It must be noted that the Jtot value used
for 40u before TDC is increased from 3 to 4 for the
same Itot value (of 41) as 30u before TDC to account
for the higher number of potential ignition centres
for the earlier BOI.
Figures 4(d) and (e) show model predictions for
the early BOIs of 50u before TDC and 60u before
TDC, which were obtained with Jtot values of 7 and
12 respectively. As mentioned before, one important
objective of the simulation is to predict combustion
at early BOIs, which yielded the best NOx–efficiency
trade-offs in the ALPING experiments. For 50u before
TDC, both the heat release rates and the cylinder
Fig. 5 Simulated results of packet equivalence ratio pressures are predicted very well. A slight delay in
distributions at the SOC for different BOIs. Note the occurrence of the peak heat release rate may be
that, in this figure, a packet equivalence ratio range observed together with a slight underprediction of
of 0.6 (for example) represents the range 0.6–0.7 the heat release duration. The heat release predic-
tion for 60u before TDC is even better. The model
As is evident from Fig. 5, for the 20u before TDC BOI, predicts the SOC, heat release duration, initial heat
all packets are richer than stoichiometric at the SOC, release rates, and peak heat release rates very well,
with nearly 50 per cent of the packets richer than the with the only downside being a slight overprediction
equivalence ratio of 2.0. Therefore, they exhibit slower of the peak pressure for 60u before TDC. Overall, the
heat release rates and, consequently, attain lower peak predictions at early BOIs are much better than for
temperatures than for intermediate BOIs. Cooler pack- late and intermediate BOIs, thus satisfying the main
ets reduce the effective flame area and lead to slower objective of the model development effort. It is also
flame combustion, thus resulting in lower peak heat interesting to note that the experimental heat release
release rates. However, the overall heat release duration profiles for the 50u before TDC and 60u before TDC
is predicted fairly well despite the overprediction of BOIs do not exhibit the initial heat release peak that
heat release rates towards the EOC for 20u before TDC. was visible for late and intermediate BOIs. Similarly,
As shown in Figs 4(b) and (c), the overall heat the heat release spikes that were noted in the
release rates are predicted better for the 30u before predicted heat release curves for late and intermedi-
TDC and 40u before TDC BOIs than for the 20u ate BOIs are virtually absent at 50u before TDC and
before TDC BOI but the peak pressures are slightly 60u before TDC. For early BOIs, diesel has sufficient
overpredicted. The predicted heat release durations time to mix with the surrounding lean natural-gas–
matched well the experimental durations for both air mixture during the long ignition delay periods,
30u before TDC and 40u before TDC. For 30u before and therefore it is possible that, upon ignition, pilot
TDC, the peak pressures are overpredicted because diesel combustion occurs simultaneously with nat-
the initial heat release rates are overpredicted. ural-gas combustion by localized flame propagation.
Compared with 20u before TDC, the packet equiva- Consequently, there is no distinct initial heat release
lence ratios for 30u before TDC are closer to peak in the experimental heat release curves. In the
stoichiometric (see Fig. 5) and so the packet tem- simulation, the long ignition delay periods for 50u
peratures are much higher, leading to faster com- before TDC and 60u before TDC resulted in packet
bustion rates. However, the peak heat release rates combustion at substantially leaner equivalence
are still slightly underpredicted. On the other hand, ratios (see Fig. 5), thus smoothing out the initial
for 40u before TDC, both the cylinder pressure and heat release spikes that were observed at late and
the heat release rate predictions are better. The intermediate BOIs. As observed in Fig. 5, a majority
peak heat release rates are predicted satisfactorily of packets for the 50u before TDC BOI and virtually
and the initial heat release rates are only slightly all packets for 60u before TDC were leaner than

JAUTO1472 Proc. IMechE Vol. 224 Part D: J. Automobile Engineering


1610 S R Krishnan and K K Srinivasan

stoichiometric at the SOC, with almost 60 per cent of


packets lying in the 0.6–0.7 equivalence ratio range
at the SOC for 60u before TDC. These lean packets
led to lower peak packet temperatures at early BOIs,
as described in the next section.

5.1.3 Peak packet temperatures


Since NOx formation is exponentially dependent on
the local temperatures in any combustion process, it
is useful to examine the packet temperatures to
understand the potential for NOx formation at
different BOIs. Flynn et al. [10] established that local
temperatures higher than approximately 2000 K will
invariably lead to unacceptable levels of engine-out
NOx emissions; therefore, it is important to keep
local temperatures below 2000 K to achieve LTC.
The simulated packet temperature trends are Fig. 7 Average peak packet temperatures (simulated)
and engine-out NOx emissions (measured) for
illustrated in Fig. 6, which shows the percentages
different BOIs
of packets that fall within a given peak packet
temperature range for different BOIs. Again, this
TpPEAK_AVG and the experimentally measured engine-
representation (in terms of percentages) is necessary
out NOx emissions at different BOIs. For each BOI, the
because different BOIs use different total numbers of
maximum (peak) temperature attained by each packet
packets. In Fig. 6, a given peak packet temperature
over its individual lifetime (from the BOI until the EOC)
range value indicates the temperature range be-
was stored in an array. For a given number of packets,
tween that value and the immediately preceding
the numerical average of all these peak packet
temperature value (lower by 100 K); for example, a
temperatures was computed and designated as the
peak packet temperature range value of 2600 K
average peak packet temperature TpPEAK_AVG. Conse-
signifies the temperature range 2500–2600 K. Fig-
quently, TpPEAK_AVG is not simply an average zone
ure 7 shows the average peak packet temperatures
temperature but the average of the peak temperatures
of packets (which were usually the hottest regions in the
cylinder, routinely exceeding 2000 K), which are prob-
ably more important for NOx formation. For reference,
the critical temperature (Tcritical NOx < 2000 K) at which
NOx formation becomes significant and the 2010 US
Environmental Protection Agency (EPA) NOx emissions
standard of 0.27 g/kW h for heavy-duty diesel engines
are also shown in Fig. 7.
In addition to the high local in-cylinder tempera-
tures, NOx formation is also dependent on the
residence times of these high-temperature regions.
To illustrate the relative importance of the residence
times of the high-temperature regions, Fig. 8 pre-
sents the peak packet temperatures and the corre-
sponding normalized packet combustion durations
(representing the in-cylinder residence times of hot
regions) for all packets at different BOIs. In this
figure, the actual combustion duration (from the
SOC until the EOC) for each ignited packet was
Fig. 6 Simulated results of the peak packet tempera- normalized with respect to the overall combustion
ture distributions for different BOIs. Note that, duration for a given BOI to obtain the normalized
in this figure, a temperature range of 2000 K (for packet combustion duration. In interpreting Fig. 8, it
example) represents the range 1900–2000 K is important to note that the peak temperatures of

Proc. IMechE Vol. 224 Part D: J. Automobile Engineering JAUTO1472


Multi-zone modelling of partially premixed low-temperature combustion 1611

engine-out NOx emissions were relatively higher


because the remaining peak packet temperatures were
distributed throughout the temperature range and
several packets were at or near stoichiometric condi-
tions at the SOC. Since more than 70 per cent of the
packets attained peak temperatures higher than 2000 K,
and about 10 per cent of the packets fell in the range
2800–2900 K, TpPEAK_AVG was much higher at 2260 K.
As observed in Fig. 7, the highest NOx emissions
were measured in the BOI range 30–40u before TDC,
indicating the highest local temperatures at these
conditions. This is confirmed in the simulated peak
temperature and normalized combustion duration
trends (see Fig. 8) for the 40u before TDC BOI, for
which nearly 60 per cent of the peak packet
temperatures fell between 2700 K and 2900 K and
their normalized combustion durations were between
0.55 and 0.9. Also, the highest TpPEAK_AVG values of
about 2700 K were obtained for the 35u before TDC
Fig. 8 Peak packet temperatures and corresponding and 40u before TDC BOIs. For the late BOI of 20u
normalized packet combustion durations (ef- before TDC, more than 98 per cent of the predicted
fective packet residence times) at different
peak packet temperatures were between 2300 K and
BOIs. Note that the combustion duration for
each packet was normalized with the overall 2600 K, leading to a much higher TpPEAK_AVG value of
combustion duration at a given BOI about 2480 K compared with the early BOI of 60u
before TDC. Virtually all packets for the 20u before
the flame zone were much lower (about 1600–1700 K) TDC BOI attained peak temperatures higher than
at all BOIs and the normalized flame residence times 2000 K and normalized combustion durations be-
were 1.0 (by definition). tween 0.35 and 1.0; the only exception was the last
From Fig. 6, it is evident that 80 per cent of injected packet in the interior regions of the spray (see
packets for the 60u before TDC BOI had peak Fig. 8), which had a very low peak temperature of
temperatures below 1900 K; this is also reflected in about 900 K and a residence time of only 0.4. The
Fig. 7 where 60u before TDC showed a TpPEAK_AVG relatively higher peak packet temperatures and the
value of 1720 K. In Fig. 8, the peak packet tempera- corresponding wide range of packet residence times
tures and the normalized combustion durations for for the 20u before TDC BOI explain the significantly
all packets at the 60u before TDC BOI are shown. For higher engine-out NOx emissions measured for 20u
this early BOI, most packets (especially the packets before TDC than for 60u before TDC. More impor-
that were injected earlier in the pilot injection tantly, these observations confirm a strong correlation
process) experienced peak temperatures signifi- between the peak local (packet) temperatures, their
cantly lower than 2000 K and relatively long normal- effective residence times (the normalized packet com-
ized combustion durations (0.7–0.95). Only a small bustion durations), and the measured NOx emissions
fraction of packets (mostly injected later in the pilot for all BOIs in ALPING combustion. In particular, if the
injection process) attained higher peak temperatures average peak local (packet) temperature exceeded the
but shorter normalized combustion durations (0.5– critical NOx formation temperature of 2000 K for any
0.6). In general, most of the packets for the 60u BOI and the packet residence times were comparable
before TDC BOI attained low peak temperatures with the overall combustion durations (normalized
(because these packets were very lean even at the durations longer than 0.5), the measured engine-out
SOC, as observed in Fig. 5) and persisted for most of NOx emissions were always higher than the 2010 US
the combustion duration. These predictions are EPA NOx standard for heavy-duty diesel engines.
consistent with the extremely low NOx emissions
measured for 60u before TDC. On the other hand,
5.1.4 Detailed results at the early BOI of 60u before TDC
even though 20 per cent of packets at 50u before
TDC had peak temperatures lower than 1900 K and Since the lowest NOx emissions were obtained at the
long normalized combustion durations (0.85–1.0), the early BOI of 60u before TDC, it is instructive to

JAUTO1472 Proc. IMechE Vol. 224 Part D: J. Automobile Engineering


1612 S R Krishnan and K K Srinivasan

examine model predictions at this BOI in greater


detail. The zonewise heat release rates and cumula-
tive heat release fractions are illustrated in Figs 9(a)
and (b) respectively. In these figures, the subscripts
p, f, and b refer to the packets, flame, and burned
zones respectively, the subscripts d and ng refer to
diesel and natural gas respectively, and the subscript
tot refers to the total heat release in the cylinder. The
cumulative heat release fractions were calculated as
ratios of cumulative heat release occurring in
different zones to the total fuel chemical energy
input from both diesel and natural gas. Most of the
combustion heat release (approximately 87 per cent)
occurred in the flame zone, with packets contribut-
ing only about 4 per cent. Figure 9(a) shows that
virtually all the packet heat release occurred early
(between 10u before TDC and 10u after top dead
centre (after TDC)) in the combustion process. The
heat release in the packets increased their tempera-
tures and led to rapid expansion of the packets,
which in turn increased the effective flame area.
Thus, the initial packet heat release rates exerted a
profound influence on flame entrainment and burn
rates through the flame area, thus affecting the flame
zone heat release rates. Since the overall combustion
heat release occurs for almost 40u CA, in contrast
with much shorter combustion durations typically
observed for HCCI-type combustion [11], ALPING
combustion is clearly slower than HCCI and results
in smoother pressure rise rates, thus enabling high-
load (BMEP, about 12 bar) engine operation. In the
packets, most of the heat release (about 3 per cent of
the total fuel energy input) is due to diesel combus-
tion and only 1 per cent arises from combustion of
entrained natural gas. This is consistent with the
actual experimental engine operating condition in
which measured diesel flow corresponded to only
about 3 per cent of the total fuel chemical energy Fig. 9 (a) Zonewise heat release rates (HRR) and (b)
input. Also, owing to the different combustion rates cumulative heat release fractions (HRcum) for
of diesel and natural gas in the packets, a delay the 60u before TDC BOI. The subscripts in this
between the onset of diesel and natural-gas heat figure refer to various zones or fuels as follows:
p, packets; p-d, packet diesel; p-ng, packet
release in the packets is observed.
natural gas; f, flame; tot, total (all zones); cum-
Figure 10 shows the temperatures of different p, cumulative packets; cum-f, cumulative
zones (the subscripts u, f, and b refer to the flame; cum-tot, cumulative total (all zones)
unburned, flame, and burned zones respectively),
the average cylinder temperature Tavg, and the peak flame temperatures (the peak Tf value is very
instantaneous maximum temperature Tinst-max much less than Tcritical NOx ) are due to the very lean
among all zones inside the cylinder. The flame and equivalence ratio (approximately 0.38) of the flame
burned zones are formed only at the SOC and, after zone. Consequently, very little NOx will be formed in
the EOC, the packets and the flame zone cease to the flame zone. At the SOC, Tavg 5 Tu but, later, Tavg
exist. Therefore, Tb and Tf are zero before the SOC increases until it becomes equal to Tb (towards the
and Tf is zero after the EOC. The peak values of Tf EOC) since most of the cylinder contents are burned.
and Tb are only about 1650 K. These relatively low Compared with Tf and Tb, the peak value of Tinst-max

Proc. IMechE Vol. 224 Part D: J. Automobile Engineering JAUTO1472


Multi-zone modelling of partially premixed low-temperature combustion 1613

evaporation and mixing) of ignition delay was much


shorter than the chemical component (pre-ignition
reactions). After the SOC, the mass fraction of diesel
burned in the packets increased rapidly and nearly 40
per cent of the injected diesel burned before TDC,
with the entire diesel being consumed before 10u
after TDC. Compared with the overall diesel burn
duration of less than 20u CA, natural-gas combustion
in the packets was much slower, persisting until
about 30u after TDC. Only 25 per cent of the total
natural gas entrained into the packets burned within
the available time; the remaining unburned natural
gas was transferred to the burned zone together with
the dumped packets. Slower burning of natural gas
(methane) in the packets is believed to occur for two
important reasons: first, the self-inhibiting nature of
methane oxidation; second, the non-inclusion of
possible enhancement of methane burn rates in the
Fig. 10 Zonewise temperature profiles for the 60u packets due to local turbulence. The former cause is
before TDC BOI. The subscripts in this figure
attributable to the chosen global kinetic scheme; i.e.
refer to various zone temperatures as follows:
u, unburned zone; f, flame zone; b, burned the negative exponent (20.3) for methane concentra-
zone; avg, average; inst-max, instantaneous tion in equations (6) and (7) inhibits reaction rates at
maximum temperature in the cylinder high methane concentrations in the packets. On the
other hand, modelling the potential enhancement of
is much higher (greater than 2700 K) and is attained methane burn rates in the packets due to local
by only a few packets (less than 2 per cent, as shown turbulence is beyond the scope of the phenomenolo-
in Fig. 5). After the initial stages of combustion (up gical framework adopted in the current simulation.
to 5u before TDC) where Tinst-max is only slightly Figure 11(b) illustrates important flame zone
higher than Tf, the hottest zones in the cylinder behaviour including the flame zone entrainment
throughout the combustion process are the packets. and burn histories, the flame temperature Tf, and the
Therefore, the simulation predicted that the diesel total number ndump of dumped packets at a given
and natural-gas combustion processes in the packets CA. The cumulative flame entrainment and burn
(as opposed to the lean flame zone) are virtually the histories clearly indicate that entrainment in the
only source of NOx formation in ALPING combustion. flame zone was faster than combustion of the
Figure 11 provides additional results, indicating entrained mass throughout the combustion process;
the important differences between the flame zone in addition, only about 95 per cent of the entrained
and combustion in the packets. Figure 11(a) shows mass burned. The flame-zone burned mass fraction
the mass fraction histories of diesel and natural gas increased rapidly after the initial combustion phase
associated with all the packets in the cylinder. All (after TDC) and exhibited a fairly linear trend
diesel fractions shown here were obtained with thereafter until near the EOC. The attainment of
respect to the total mass of diesel injected into the the EOC was triggered by the commencement of the
cylinder, while the natural-gas burned fraction was dumping process (shown by the increase in ndump)
obtained with respect to the total mass of natural gas around 25u after TDC, which led to a decrease in the
entrained in the packets throughout the combustion flame area because the dumped packets no longer
process. The entire diesel quantity is injected within contributed to the flame area computations. Owing
5u CA after the BOI. Diesel evaporation started at the to this decrease in the flame area after the start of
BOI and, by the time that injection is over, about 30 dumping and the reduced availability of fresh
per cent of diesel had evaporated. This is also natural-gas–air mixture in the unburned zone, the
reflected in the behaviour of the liquid diesel fraction flame mass entrainment as well as the burn rate
with the maximum of about 70 per cent occurring at tapered off.
the EOI. The entire diesel evaporation process for the Figure 11(c) illustrates the peak packet tempera-
60u before TDC BOI occurred within 20u CA after the tures as functions of the axial packet number I for
BOI. Consequently, the physical component (fuel different values of the radial stratification index J.

JAUTO1472 Proc. IMechE Vol. 224 Part D: J. Automobile Engineering


1614 S R Krishnan and K K Srinivasan

Fig. 11 Detailed results for the 60u before TDC BOI illustrating (a) the packet mass fractions, (b)
the flame-zone mass fractions, flame temperature, and total number of dumped
packets, and (c) the peak packet temperatures

The peak packet temperatures varied between 5.2 Parametric studies


1500 K and 2700 K over the range of I and J values.
In this section, the predictive capabilities of the
For all J values, the peak packet temperatures
phenomenological model are explored. The pre-
increased with increasing I values, which represent
dicted effects of three important engine operating
packets injected later in the injection process.
parameters including the intake charge temperature
However, the magnitude of this increase was greater
Tin, the pilot fuel quantity Qinj, and the natural-gas
for higher J values, which represent packets that
equivalence ratio wNG on ALPING combustion are
were nearer to the spray axis (e.g. J 5 12). Interior
examined for the early BOI of 60u before TDC.
packets that were also injected towards the end of
the injection process experienced less entrainment
and attained relatively higher equivalence ratios 5.2.1 Intake charge temperature effects
(closer to stoichiometric) than packets in the spray
periphery (e.g. J 5 1). Therefore, these interior pack- Figure 12 illustrates the effects of Tin on the predicted
ets experienced faster combustion rates and attained cylinder pressure, heat release rate, and peak packet
higher peak temperatures. temperature for 60u before TDC. As Tin is increased

Proc. IMechE Vol. 224 Part D: J. Automobile Engineering JAUTO1472


Multi-zone modelling of partially premixed low-temperature combustion 1615

Fig. 12 Effects of the intake charge temperature Tin on (a) the predicted cylinder pressure, (b)
the predicted heat release rate, and (c) the predicted peak packet temperature ranges for
the 60u before TDC BOI

from 75 uC to 105 uC, the onset of ignition is also advanced as Tin was increased. A direct outcome
progressively advanced, the peak cylinder pressures of faster combustion heat release was shorter com-
increase from about 90 bar to more than 120 bar, and bustion durations. The combustion duration de-
the location of the peak pressures also moves closer to creased by about 10u CA with increasing Tin: from
the TDC. The reasons for these trends can be found in more than 40u CA for 75 uC to 30u CA for 105 uC. The
the heat release rate profiles shown in Fig. 12(b). overall improvement in the combustion heat release
Clearly, the higher Tin led to faster pre-ignition rate, the shorter combustion duration, and the earlier
reactions and advanced the SOC by more than 10u combustion phasing at higher Tin are reflected in the
CA. In addition, the peak heat release rates increased fact that a greater fraction of packets reach higher
from about 225 J/deg for 75 uC to about 275 J/deg for peak temperatures (see Fig. 12(c)). For instance,
105 uC and the location of the peak heat release was while no packets attained the peak temperatures in

JAUTO1472 Proc. IMechE Vol. 224 Part D: J. Automobile Engineering


1616 S R Krishnan and K K Srinivasan

the 2800–2900 K range for 75 uC, about 7 per cent of 5.2.2 Pilot fuel quantity effects
the peak packet temperatures fell in this range for
105 uC. On the other hand, as Tin was increased from The effects of Qinj on the predicted cylinder pressure,
75 uC to 105 uC, the fraction of packets with peak heat release rate, and peak packet temperature are
temperatures lower than 1900 K decreased from 80 shown in Fig. 13. As Qinj is increased from 3.3 g/min
per cent to less than 30 per cent. Also, the average to 5.4 g/min, ignition occurs earlier, the peak
peak packet temperatures TpPEAK_AVG increased from cylinder pressures increase, and the location of the
1720 K at 75 uC to 2150 K at 105 uC, which is greater peak pressures occurs closer to the TDC. The heat
than the Tcritical NOx value of 2000 K. These trends in release rate profiles (see Fig. 13(b)) show that higher
the local (packet) temperatures indicate the potential Qinj led to a slight advancement of the SOC by about
for higher NOx formation at higher Tin. 2u CA, substantially higher peak heat release rates

Fig. 13 Effects of the pilot fuel quantity Qinj on (a) the predicted cylinder pressure, (b) the
predicted heat release rate, and (c) the predicted peak packet temperature ranges for the
60u before TDC BOI

Proc. IMechE Vol. 224 Part D: J. Automobile Engineering JAUTO1472


Multi-zone modelling of partially premixed low-temperature combustion 1617

(from about 225 J/deg for 3.3 g/min to about 375 J/ ranges is relatively unaffected by wNG with the
deg for 5.4 g/min), and earlier peak heat release exceptions of the temperature ranges below 1900 K
phasing. Higher Qinj resulted in hotter packets that and 2700–2800 K. As wNG was increased from 0.37 to
expanded rapidly, increasing the effective flame area 0.56, the fraction of low-peak-temperature (less than
and overall combustion rates. Compared with the 1900 K) packets decreased from 80 per cent to less
effects of Tin, the SOC advancement was weaker but than 70 per cent while the fraction of high-peak-
the peak heat release rates were much higher at temperature packets increased from 2 per cent to 10
higher Qinj. The higher heat release rates resulted in per cent in the 2700–2800 K range. On the other hand,
shorter combustion durations (from 40u CA for 3.3 g/ there was only a small increase in TpPEAK_AVG from
min to 25u CA for 5.4 g/min). All these combustion 1720 K at wNG 5 0.37 to 1850 K at wNG 5 0.56. While the
effects at higher Qinj combined to cause more higher TpPEAK_AVG may lead to slightly higher NOx
packets to reach higher peak temperatures, as emissions at higher wNG, this increase will probably be
evident from Fig. 13(c). As Qinj was increased from modest compared with the influences of Tin and Qinj
3.3 g/min to 5.4 g/min, the fraction of low-tempera- on NOx. This is largely because, as wNG was increased
ture packets (with peak temperatures lower than to 0.56, the TpPEAK_AVG value attained was still lower
1900 K) decreased from 80 per cent to only 10 per than the critical NOx formation temperature of
cent while high-temperature packets in the 2800– 2000 K. Consequently, engine-out NOx emissions
2900 K range increased from 0 per cent to 14 per with ALPING combustion remained very low even
cent. With increasing Qinj, TpPEAK_AVG increased from when the engine load was increased by increasing
1720 K at 3.3 g/min to 2270 K at 5.4 g/min. These wNG, a fact that was demonstrated experimentally in
trends confirm conventional wisdom in dual-fuel previous studies [18, 20].
combustion, which dictates that a higher Qinj will
probably lead to higher NOx emissions owing to
higher local temperatures. Further, when Figs 12 and 6 CONCLUSIONS
13 are compared, the impact of increasing Qinj on
the local temperatures (and the NOx emissions) is
A multi-zone phenomenological simulation has been
clearly more significant than increasing Tin.
developed to simulate partially premixed ALPING
LTC. This simulation combined, in a novel way,
5.2.3 Natural-gas equivalence ratio effects several modelling aspects of diesel spray combustion
and premixed turbulent flame propagation. The
Figure 14 demonstrates the combustion behaviour cylinder contents were divided into an unburned
for the 60u before TDC BOI at different wNG but fixed zone, pilot fuel zones (or ‘packets’) that account for
Qinj (5 3.3 g/min) and Tin (5 75 uC). Since all the other diesel combustion, a flame zone for natural-gas
parameters were fixed, increasing wNG from 0.37 to combustion, and a burned zone. Predictions from
0.56 led to an overall increase (approximately 40 per the simulation were validated against experimental
cent) in the engine load (IMEP). The cylinder pressure results and parametric studies were performed,
results shown in Fig. 14(a) followed expectations in leading to the following significant conclusions.
that the peak pressures increased as the natural-gas
equivalence ratio was increased from 0.37 to 0.56. 1. The simulation predicted ALPING LTC satisfacto-
Figures 14(b) and (c) show the heat release profiles rily over a wide range of pilot injection timings
and peak packet temperatures respectively at differ- (BOI) from 20u before TDC to 60u before TDC.
ent wNG. First, with increasing wNG, there is a slight 2. The onset of ignition (predicted with the Shell
(although noticeable) delay in the SOC. Second, the autoignition model) and heat release profiles
peak heat release rates are significantly higher and the (especially combustion durations) were predicted
combustion durations are slightly shorter at higher accurately for the entire BOI range but the peak
wNG. Third, the initial heat release rates (within the pressures were slightly overpredicted for the 30u,
first 10u CA after the SOC) are not affected much by 40u, and 50u before TDC BOIs. Both the combus-
increasing wNG since the initial combustion rates in tion pressure and the heat release profiles were
the packets were probably similar as Tin and Qinj were predicted better at early BOIs, which yielded very
fixed at 75 uC and 3.3 g/min respectively. The influ- low NOx emissions in ALPING combustion ex-
ence of wNG is noticeable only later (after TDC) during periments.
flame-zone combustion. Finally, the distribution of 3. Strong coupling was observed between pilot spray
the peak packet temperatures in most temperature combustion in the packets and premixed turbu-

JAUTO1472 Proc. IMechE Vol. 224 Part D: J. Automobile Engineering


1618 S R Krishnan and K K Srinivasan

Fig. 14 Effects of the natural-gas equivalence ratio wNG on (a) the predicted cylinder pressure,
(b) the predicted heat release rate, and (c) the predicted peak packet temperature ranges
for the 60u before TDC BOI

lent combustion in the flame zone because pilot measured engine-out NOx emissions were extre-
diesel combustion in the packets directly affected mely low at the 60u before TDC BOI, probably
the flame area and, consequently, flame-zone because of relatively low-temperature packets that
combustion. Further, it was shown that the persisted longer in the combustion process;
number of ignition centres (packets) had a further, the average peak packet temperature
profound influence on the overall heat release TpPEAK_AVG was only 1720 K, much lower than the
rates and flame-zone combustion. critical NOx formation temperature Tcritical NOx of
4. The highest in-cylinder temperatures were ob- 2000 K. By comparison, at 20u before TDC and 40u
tained in packets, indicating that pilot diesel before TDC, the peak temperatures of long-
spray combustion is probably the dominant source residence-time packets were much higher (the
of NOx emissions in ALPING combustion. The TpPEAK_AVG values for 20u before TDC and 40u

Proc. IMechE Vol. 224 Part D: J. Automobile Engineering JAUTO1472


Multi-zone modelling of partially premixed low-temperature combustion 1619

before TDC were 2480 K and 2700 K respectively), ACKNOWLEDGEMENTS


resulting in substantially higher NOx emissions.
5. Although approximately 90 per cent of the The authors appreciate the partial financial support
combustion energy release occurred in the flame for this work provided by the Sustainable Energy
zone, the predicted peak flame temperature was Research Center at Mississippi State University. In
only about 1650 K (owing to the very lean flame- addition, the authors gratefully acknowledge the
zone equivalence ratio of approximately 0.38), detailed comments and suggestions for improve-
indicating the near-zero potential for NOx forma- ment provided by the two anonymous reviewers.
tion in the flame zone.
6. Increasing the intake charge temperature Tin and F Authors 2010
the pilot fuel quantity Qinj led to faster combus-
tion, higher peak heat release rates, and higher REFERENCES
TpPEAK_AVG values at the 60u before TDC BOI,
confirming their relative importance vis-à-vis 1 Zhang, F., Okamoto, K., Morimoto, S., and Shoji,
NOx emissions. However, increasing the natural- F. Methods of increasing the BMEP (power output)
gas equivalence ratios wNG at fixed Tin and Qinj led for natural-gas SI engines. SAE paper 981385, 1998.
to only higher peak heat release rates without 2 Kopecek, H., Wintner, E., Lackner, M., Winter, F.,
affecting the TpPEAK_AVG values significantly. and Hultqvist, A. Laser-stimulated ignition in a
homogeneous charge compression ignition engine.
These results are in agreement with experimen-
SAE paper 2004-01-0937, 2004.
tally measured trends [18, 20] for NOx emissions, 3 Bihari, B., Gupta, S. B., Sekar, R. R., Gingrich, J.,
which were relatively unaffected as wNG was and Smith, J. Development of advanced laser
increased (at fixed Qinj) to attain engine loads as ignition system for stationary natural gas recipro-
high as 12 bar BMEP. cating engines. In Proceedings of the Fall Technical
Meeting of the Internal Combustion Engines
Division of ASME, Ottawa, Ontario, Canada, 11–
14 September 2005, paper ICES2005-1325 (ASME
7 FUTURE WORK International, New York).
4 Gebert, K., Beck, N. J., Barkhimer, R. L., and
In the present work, a phenomenological framework Wong, H.-C. Strategies to improve combustion and
based on pilot diesel spray combustion and pre- emission characteristics of dual-fuel pilot ignited
mixed turbulent flame propagation was developed to natural gas engines. SAE paper 971712, 1997.
5 Workman, J. and Beshouri, G. M. Single cylinder
simulate partially premixed ALPING LTC. The
testing of a high pressure electronic fuel injector for
simulation integrated pilot spray combustion and low NOx emission dual fuel engines. Part II:
premixed turbulent flame combustion through a optimization, startability and inflammability test-
unique flame area computation methodology. While ing. Trans. ASME, J. Engng Gas Turbines Power,
the present simulation predicted the onset of 1990, 112, 422–430.
ignition and heat release trends satisfactorily and 6 Karim, G. A. Combustion in gas fueled compres-
revealed several important features of ALPING LTC sion: ignition engines of the dual fuel type. Trans.
(including strong coupling between spray combus- ASME, J. Engng Gas Turbines Power, 2003, 125,
tion and flame propagation), there is room for 827–836.
7 Krishnan, S. R., Biruduganti, M., Mo, Y., Bell, S.
further improvement. In particular, the spray model
R., and Midkiff, K. C. Performance and heat release
can be improved to capture the radial stratification analysis of a pilot-ignited natural gas engine. Int. J.
better (instead of using different Jtot values at Engine Res., 2002, 3(3), 171–184. DOI: 10.1243/
different BOIs), to incorporate a more sophisticated 14680870260189280.
spray penetration model (see, for example, reference 8 Galal, M. G., Abdel Aal, M. M., and El Kady, M. A. A
[34]), and to accommodate possible spray wall comparative study between diesel and dual-fuel
impingement at early BOIs. Also, the turbulent engines: performance and emissions. Combust. Sci.
flame propagation model and the flame area Technol., 2002, 174, 241–256.
computation methodology can be refined to elim- 9 Gebert, K., Beck, N. J., Barkhimer, R. L., Wong,
H.-C., and Wells, A. D. Development of pilot fuel
inate or reduce the dependence of the combustion
injection system for CNG engine. SAE paper
rate on the number of packets, and to account for 961100, 1996.
possible spatial recombination of enflamed packets, 10 Flynn, P. F., Hunter, G. L., Farrell, L., Durrett, P.,
which may reduce the effective flame area during Akinyemi, O., Zur Loye, A. O., Westbrook, C. K.,
combustion. and Pitz, W. J. The inevitability of engine-out NOx

JAUTO1472 Proc. IMechE Vol. 224 Part D: J. Automobile Engineering


1620 S R Krishnan and K K Srinivasan

emissions from spark ignited and diesel engines. gas engine. Int. J. Engine Res., 2007, 8(3), 289–303.
Proc. Combust. Inst., 2000, 28, 1211–1218. DOI: 10.1243/14680874JER02306.
11 Zhao, F., Asmus, T., Assanis, D., Dec, J. E., Eng, J. 24 Jung, D. and Assanis, D. N. Multi-zone DI diesel
A., and Najt, P. M. Homogeneous charge compres- spray combustion model for cycle simulation
sion ignition (HCCI) engines, 2003 (SAE Interna- studies of engine performance and emissions.
tional, Warrendale, Pennsylvania). SAE paper 2001-01-1246, 2001.
12 Hasegawa, R. and Yanagihara, H. HCCI combus- 25 Yoshizaki, T., Nishida, K., and Hiroyasu, H.
tion in DI diesel engine. SAE paper 2003-01-0745, Approach to low NOx and smoke emission engines
2003. by using phenomenological simulation. SAE paper
13 Kimura, S., Aoki, O., Kitahara, Y., and Aiyoshi- 930612, 1993.
zawa, E. Ultra-clean combustion technology com- 26 Patriotis, E. G. and Hountalas, D. T. Validation of a
bining a low-temperature and premixed combus- newly-developed quasi-dimensional combustion
tion concept for meeting future emission stan- model – application on a heavy-duty DI diesel
dards. SAE paper 2001-01-0200, 2001. engine. SAE paper 2004-01-0923, 2004.
14 Kalghatgi, G., Risberg, P., and Angstrom, H.-E. 27 Hountalas, D. T. and Papagiannakis, R. G. A
Advantages of fuels with high resistance to auto- simulation model for the combustion process of
ignition in late-injection, low-temperature, com- natural gas engines with pilot diesel fuel as an
pression ignition combustion. SAE paper 2006-01- ignition source. SAE paper 2001-01-1245, 2001.
3385, 2006. 28 Pirouzpanah, V. and Kashani, B. O. Prediction of
15 Kalghatgi, G. Auto-ignition quality of practical major pollutants emission in direct-injection, dual-
fuels and implications for fuel requirements of fuel diesel and natural gas engines. SAE paper
future SI and HCCI engines. SAE paper 2005-01- 1999-01-0841, 1999.
0239, 2005. 29 Liu, Z. and Karim, G. A. A predictive model for the
16 Risberg, P., Kalghatgi, G., Angstrom, H.-E., and combustion process in dual fuel engines. SAE
Wahlin, F. Auto-ignition quality of diesel-like fuels paper 952435, 1995.
in HCCI engines. SAE paper 2005-01-2127, 2005. 30 Liu, Z. and Karim, G. A. Simulation of combustion
17 Musculus, M. P. B. Multiple simultaneous optical processes in gas-fuelled diesel engines. Proc. IM-
diagnostic imaging of early-injection low-tempera- echE, Part A: J. Power and Energy, 1997, 211(2),
ture combustion in a heavy-duty diesel engine. SAE 159–169. DOI: 10.1243/0957650971537079.
paper 2006-01-0079, 2006. 31 Khalil, E. B. and Karim, G. A. A kinetic investiga-
18 Krishnan, S. R., Srinivasan, K. K., Singh, S., Bell, tion of the role of changes in the composition of
S. R., Midkiff, K. C., Gong, W., Fiveland, S. B., and natural gas in engine applications. Trans. ASME, J.
Willi, M. Strategies for reduced NOx emissions in Engng Gas Turbines, 1997, 124, 404–411.
pilot ignited natural gas engines. Trans. ASME, J. 32 Kusaka, J., Tsuzuki, K., Daisho, Y., and Saito, T. A
Engng Gas Turbines Power, 2004, 126, 665–671. numerical study on combustion and exhaust gas
19 Singh, S., Krishnan, S. R., Srinivasan, K. K., emissions characteristics of a dual-fuel natural gas
Midkiff, K. C., and Bell, S. R. Effect of pilot engine using a multi-dimensional model combined
injection timing, pilot quantity, and intake charge with detailed kinetics. SAE paper 2002-01-1750,
conditions on performance and NOx emissions for 2002.
an advanced low-pilot-ignited natural gas engine. 33 Dent, J. C. Basis for the comparison of various
Int. J. Engine Res., 2004, 5(4), 329–348. DOI: 10.1243/ experimental methods for studying spray penetra-
146808704323224231. tion. SAE paper 710571, 1971.
20 Srinivasan, K. K., Krishnan, S. R., Singh, S., 34 Siebers, D. L. Scaling liquid-phase fuel penetration
Midkiff, K. C., Bell, S. R., Gong, W., Fiveland, S. in diesel sprays based on mixing-limited vaporiza-
B., and Willi, M. The advanced low pilot ignited tion. SAE paper 1999-01-0528, 1999.
natural gas engine – a combustion analysis. Trans. 35 Bell, S. R. Development of a cycle simulation for a
ASME, J. Engng Gas Turbines Power, 2006, 128, coal-fueled direct-injected, internal combustion en-
213–218. gine. Doctoral dissertation, Texas A&M University,
21 Egolfopoulos, F. N., Holley, A. T., and Law, C. K. College Station, Texas, USA, 1986.
An assessment of the lean flammability of CH4/air 36 Kanury, A. Introduction to combustion phenom-
and C3H8/air mixtures at engine-like conditions. ena, 1975 (Gordon and Breach, New York).
Proc. Combust. Inst., 2007, 31, 3015–3022. 37 Krishnan, S. R., Srinivasan, K. K., and Midkiff, K. C.
22 Srinivasan, K. K., Krishnan, S. R., Qi, Y., Yang, H., Ignition in pilot-ignited natural gas low temperature
and Midkiff, K. C. Analysis of diesel pilot-ignited combustion: multi-zone modeling and experimen-
natural gas low-temperature combustion with hot tal results. In Proceedings of the ASME Internal
exhaust gas recirculation. Combust. Sci. Technol., Combustion Engine Division 2009 Spring Technical
2007, 179(9), 1737–1776. Conference, Milwaukee, Wisconsin, USA, 3–6 May
23 Qi, Y., Srinivasan, K. K., Krishnan, S. R., Yang, H., 2009, paper ICES 2009-76145 (ASME International,
and Midkiff, K. C. Effect of hot exhaust gas New York).
recirculation on the performance and emissions 38 Krishnan, S. R. Experimental investigations and
of an advanced injection low pilot-ignited natural phenomenological simulation of combustion in a

Proc. IMechE Vol. 224 Part D: J. Automobile Engineering JAUTO1472


Multi-zone modelling of partially premixed low-temperature combustion 1621

low pilot-ignited natural gas engine with a focus on 54 Hountalas, D. T., Kouremenos, D. A., and Five-
advanced injection timings. Doctoral dissertation, land, S. B. Some considerations on the estimation
University of Alabama, Tuscaloosa, Alabama, USA, of the heat release of DI diesel engines using
2005. modeling techniques. SAE paper 2004-01-1405,
39 Halstead, M. P., Kirsch, L. J., and Quinn, C. P. The 2004.
autoignition of hydrocarbon fuels at high tempera-
tures and pressures – fitting of a mathematical
model. Combust. Flame, 1977, 30, 45–60. APPENDIX
40 Sazhina, E. M., Sazhin, S. S., Heikal, M. R., and
Marooney, C. J. The Shell autoignition model: Notation
applications to gasoline and diesel fuels. Fuel, 1999,
78(4), 389–401. ALPING advanced (injection) low-pilot-
41 Westbrook, C. K. and Dryer, F. L. Simplified ignited natural gas
reaction mechanisms for the oxidation of hydro- after TDC after top dead centre
carbon fuels in flames. Combust. Sci. Technol.,
Af flame area
1981, 27, 31–43.
42 Heywood, J. B. Combustion and its modeling in Ap3 pre-exponential factor controlling
spark-ignition engines. In Proceedings of the Third fuel consumption in the propagation
International Symposium on Diagnostics and mod- cycle of the Shell model
eling of combustion in internal combustion en- BMEP brake mean effective pressure
ngines (COMODIA 1994), Yokohama, Japan, 11–14 BOI beginning of (pilot) injection
July 1994, pp. 1–15 (Engine Systems Division, Japan before TDC before top dead centre
Society of Mechanical Engineers, Tokyo).
CA crank angle
43 Tabaczynski, R. J., Ferguson, C. R., and Radhak-
rishnan, K. A turbulent entrainment model for cb burned mass fraction exponent in the
spark-ignition engine combustion. Trans. SAE, flame area equation
1977, 86, 2414–2433; also SAE paper 770647, 1977. cfdrop model constant in the exponential
44 Tennekes, H. Simple model for the small-scale term for flame entrainment
structure of turbulence. Phys. Fluids, 1968, 11(3), cv specific heat at constant volume
669–671. Cw lean packet equivalence ratio
45 Tabaczynski, R. J., Trinker, F. H., and Shannon, B. exponent
A. S. Further refinement and validation of a tur-
bulent flame propagation model for spark-ignition EGR exhaust gas recirculation
engines. Combust. Flame, 1980, 39(2), 111–121. EOC end of combustion
46 Krishnan, S. R. Heat release analysis of dual fuel EOI end of injection
combustion in a direct injection compression-igni- EPA (US) Environmental Protection
tion engine. MS Thesis, University of Alabama, Agency
Tuscaloosa, Alabama, USA, 2001. EA activation energy
47 Turns, S. R. An introduction to combustion:
HC (unburned) hydrocarbon
concepts and applications, 2nd edition, 2001
(McGraw-Hill, New York). HCCI homogeneous charge compression
48 Brehob, D. D. and Newman, C. E. Monte Carlo ignition
simulation of cycle by cycle variability. SAE paper h specific enthalpy
922165, 1992. IMEP indicated mean effective pressure
49 Shen, H., Hinze, P. C., and Heywood, J. B. A study IVC intake valve closure
of cycle-to-cycle variations in SI engines using a I packet classification based on their
modified quasi-dimensional model. SAE paper
time of entry into the cylinder
961187, 1996.
50 Hill, P. G. Cyclic variations and turbulence struc-
J packet classification accounting for
ture in spark-ignition engines. Combust. Flame, radial stratification in the spray
1988, 72, 73–89. K entrainment constant
51 Tennekes, H. and Lumley, J. L. A first course in KAf flame area proportionality constant
turbulence, 1972 (MIT Press, Cambridge, Massa- LTC low-temperature combustion
chusetts). L integral length scale
52 Heywood, J. B. Internal combustion engine funda- m mass of different zones
mentals, 1988, p. 143 (McGraw-Hill, New York).
ṁ reaction rate or rate of change in the
53 Woschni, G. A universally applicable equation for
the instantaneous heat transfer coefficient in the mass
internal combustion engine. Trans. SAE, 1967, 76, MW molecular weight
3065–3083; also SAE paper 670931, 1967. NOx oxides of nitrogen

JAUTO1472 Proc. IMechE Vol. 224 Part D: J. Automobile Engineering


1622 S R Krishnan and K K Srinivasan

ndump number of packets that have been m dynamic viscosity


dumped n kinematic viscosity
PM particulate matter r density
P instantaneous cylinder pressure tb characteristic burn time
Q heat release, heat transfer, or pilot w equivalence ratio
quantity (depending on the v crankshaft angular velocity
subscript)
RON research octane number
R characteristic gas constant Subscripts
Rf equivalent flame radius calculated
avg average (over the entire cylinder)
from flame area
b burned zone or burn-related
SMD Sauter mean diameter
parameter
SOC start of combustion
ch characteristic
SL laminar burning velocity
d diesel
TDC top dead centre
T temperature e, ent entrained
Tcritical critical temperature at which NOx expt experimental
NOx
formation becomes significant f flame zone
(about 2000 K) i initial value
TpPEAK_AVG average peak packet temperature in intake manifold
u turbulent entrainment velocity inj injected
u9 turbulence intensity i zone index
V instantaneous cylinder volume ij packet (I, J)
ypb packet burned mass fraction ng, NG natural gas
Y ratio of time elapsed since the BOI to p packets (pilot fuel zones)
total injection duration sim simulated
tot overall, total, or for the entire
contents in the cylinder
g Kolmogorov microscale tot-ent total entrained
h crank angle u unburned zone
l Taylor scale wall cylinder wall

Proc. IMechE Vol. 224 Part D: J. Automobile Engineering JAUTO1472

Anda mungkin juga menyukai